Fragaria

advertisement
Multiplexed Fragaria Chloroplast Genome Sequencing
W. Njuguna
Department of Horticulture
Oregon State University
Corvallis, Oregon 97331
USA
A. Liston
Department of Botany and Plant Pathology
Oregon State University
Corvallis, Oregon 97331
USA
R. Cronn
United States Forest Service
Pacific Northwest Research Station
Corvallis, Oregon, 97331
USA
N.V. Bassil
USDA/ARS
National Clonal Germplasm Repository
Corvallis, Oregon 97333
USA
Keywords: high-throughput sequencing, polyploidy, reference genome, sequence
alignments, microreads
Abstract
A method to sequence multiple chloroplast genomes using ultra high
throughput sequencing technologies was recently described. Complete chloroplast
genome sequences can resolve phylogenetic relationships at low taxonomic levels and
identify informative point mutations and indels. The objective of this research was to
sequence multiple Fragaria chloroplast genomes using the Illumina Genome
Analyzer. Sixty-three PCR fragments from 22 species were sequenced in four
multiplex sequencing runs. Plastome sequences were assembled using a combination
of de novo and reference guided assembly with F. vesca ‘Hawaii 4’ providing the
reference. De novo assembly resulted in plastome coverage of 43-74%, and reference
guided assembly increased the plastome coverage to 76-82%. The alignment of
sequenced chloroplast genomes was 96,438 bp and contained 319 parsimonyinformative sites which will be useful in identifying chloroplast genome regions of
high sequence variation for testing evolutionary relationships.
INTRODUCTION
Strawberry, Fragaria L., is in the Rosoideae subfamily in Rosaceae (Potter et al.,
2007). The genus Fragaria was classified in the Fragariinae subtribe, Potentilleae tribe in
the Rosodae superfamily. Fragaria species include thirteen diploids (2n=2x=14), five
tetraploids (2n=4x=28), one hexaploid (2n=6x=42), four octoploids (2n=8x=56) (Staudt,
2008) and one decaploid (2n=10x=70) (Hummer et al., 2009). F. iturupensis, whose
octoploid forms were previously reported (Staudt, 1989; Staudt and Olbricht, 2008), is
also represented by decaploid forms (Hummer et al., 2009), which constituted the first
reported naturally occurring decaploid strawberry species. Resolution of Fragaria
phylogeny is not only useful for identification of sources of useful genes (Bringhurst and
Voth, 1984; Lawrence et al., 1990), but also for verification of current species
designations by inference from predicted species relationships. The need to verify
Fragaria species designations became evident following flow cytometry (Hummer and
Bassil, 2008; Hummer et al., 2009), simple sequence repeat (SSR) data analysis (Njuguna
and Bassil, 2008), and species introductions and revisions to the strawberry germplasm
collection at the USDA/ARS/NCGR in Corvallis, Oregon by strawberry expert Günter
Staudt (1928-2008).
In Fragaria, phylogenetic analyses have used chloroplast (Harrison et al., 1997)
and nuclear genome sequences (Potter et al., 2000; Rousseau-Gueutin et al., 2009), but
relationships remain unclear. The suggested Fragaria octoploid genome models
AAA’A’BBB’B’ (Bringhurst, 1990) and YYY’Y’ZZZZ/YYYYZZZZ (Rousseau-Gueutin
et al., 2009), suggest the contribution of two to four diploids to the octoploid genome. The
diploid donor species are still not known, but evidence based on grouping of octoploid
Proc. IS on Molecular Markers in Horticulture
Eds.: N.V. Bassil and R. Martin
Acta Hort. 859, ISHS 2010
315
and diploid nuclear genes (ADH, alcohol dehydrogenase; GBSSI-2 or Waxy; and DHAR,
dehydroascorbate reductase) points to F. vesca L., F. mandschurica Staudt sp. nova and
F. iinumae Makino (Davis and DiMeglio, 2004; Harrison et al., 1997; Potter et al., 2000;
Rousseau-Gueutin et al., 2009; Senanayake and Bringhurst, 1967) as possible
contributors. Based on low copy nuclear gene sequences, Rousseau-Gueutin et al. (2009)
grouped Fragaria diploids into three clades. However, the diploid species within clade X
were poorly resolved and the placement of F. bucharica remains unclear. The use of
nuclear genes for phylogenetic analysis is complicated by polyploidy and recombination
(Nishikawa et al., 2002), making the chloroplast genome an attractive option for
Fragaria.
Further exploitation of the chloroplast genome in Fragaria for phylogenetic
analysis is warranted due its small size, non-recombinant nature, and high sequence
conservation, all factors that reduce the complexity of analysis and interpretation of
results. For efficient use of limited chloroplast sequence divergence, large-scale
sequencing is required, and is now possible with high throughput sequencing platforms.
Sequencing of multiple small genomes to high coverage depth using high-throughput
sequencing platforms was recently demonstrated (Cronn et al., 2008). In this study, we
used multiplex sequencing to uncover sequence divergences in the chloroplast genomes
of Fragaria species.
MATERIALS AND METHODS
Plant Material and DNA Extraction
Twenty-one of 22 wild Fragaria species, and one Potentilla villosa accession, a
close relative of Fragaria in the Rosaceae family, that have been preserved at the
germplasm repository in Corvallis (Table 1) were included in the study. DNA was
extracted from actively-growing leaves using a modified protocol based on the
PUREGENE® kit (Gentra Systems Inc. Minneapolis, MN). The DNA samples were
cleaned further with Nucleon PhytoPureTM resin, a component found in the illustra
Nucleon PhytopureTM kits for plant and fungal DNA extraction (GE Healthcare UK
Limited, Buckinghamshire, UK). DNA quality and quantity was analyzed with a Wallac
1420 VictorV microplate reader (PerkinElmer, Waltham, MA). DNA concentration was
adjusted to 3 ng/µl for PCR.
Primer Selection, Design and PCR
A total of 203 chloroplast primers (108 forward, 95 reverse) were screened in
various logical combinations in four species to identify primer pairs that amplify >2500
bp fragments and provide maximum coverage of the chloroplast genome. Where possible,
primer pairs that amplified single bands and that amplified in most or all of the species
were chosen. Of the 203 primers, 141 had been used to amplify the Cucumis sativus L.
chloroplast genome (Chung et al., 2007), 25 were designed from the complete genome
sequence of Morus indica ‘K2’ (Ravi et al., 2006), and 36 were designed from the nearly
complete chloroplast genome of F. vesca cv. Hawaii4 (http://strawberry.vbi.vt.edu/tiki­
index.php). Sixty-three primer pairs were chosen to amplify the entire chloroplast
genomes of 21 Fragaria species and one Potentilla accession (list available upon request
from corresponding author). Long-range PCR was carried out using PhusionTM HighFidelity DNA polymerase (New England Biolabs, Ipswich, MA). Amplifications were
performed in 10 µl total reaction volumes containing 5x Phusion GC buffer, 2.5 mM of
each dNTP, 10 µM of each primer, 5 U of Phusion DNA polymerase, 0.05 µl of 3%
DMSO and 5 ng of DNA template. PCR product quantification was carried out using the
Quant-iTTM PicoGreen® dsDNA quantification protocol (Molecular Probes, Inc. Eugene,
OR) manufacturer’s specifications. Equimolar amounts of PCR products were pooled for
each species to generate 1-5 µg of chloroplast DNA which was prepared for sequencing
using the Illumina sample preparation kit (Illumina Inc., San Diego, CA) and the
modifications of Cronn et al. (2008). Each PCR pool was sheared using nitrogen from a
316
compressed source at 42 psi for two minutes. The fragments were repaired to remove 3’
and fill 5’ overhangs before adding ‘A’ bases to the 3’ ends to allow the ligation of single
‘T’ bases on the 3’ end of the adapters. Illumina adapters modified by the addition of a
three base pair barcode were ligated to the fragments; a different barcode was used for
each species in a multiplex pool. DNA fragments of approximately 300 bp were isolated
by cutting the fragments from a 2% low melting agarose gel and PCR was performed to
enrich for the targeted fragment sizes. Fragments from five or six chloroplast genomes
were then mixed in equimolar ratios at a final concentration of 10 nM per pool in each of
four multiplex pools.
Sequence Data Analysis
After the sequencing run, raw image data for each sequencing cycle was processed
into base calls and alignment files through the Illumina/Solexa Pipeline (version 0.2.2.6).
Binning was carried out using the three base pair nucleotide tags. After sorting
microreads into species-specific bins, the barcodes and adapter tags were removed and
resulting 32 bp microreads used for subsequent analysis. Contigs of F. vesca ‘Baron
Solemacher’ were assembled de novo using Velvet Assembler 0.4 (Zerbino and Birney,
2008), a hash length of 17, minimum coverage 5x, and minimum contig length of 100.
The contigs were then aligned to the 135,946 bp chloroplast genome of F. vesca
‘Hawaii4’. The alignment was manually checked for errors, with insertions and deletions
for a more accurate alignment. Non-target sequences from other cellular genomes were
discarded and a consensus sequence was generated in Bioedit. The ‘Baron Solemacher’
sequence was 132,287 bp in length and served as the reference for subsequent analyses.
The reference genome was labeled ‘FvRefv2’ and annotated using DOGMA
(http://dogma.ccbb.utexas.edu/).
Contigs for each of the 22 species were assembled as described for F. vesca
‘Baron Solemacher’. ‘N’s were added to the ends of the contigs from each of the 22
species to aid in distinguishing gaps in the sequence from indels in the alignment. The Nended contigs were then aligned to the FvRefv2 reference using CodonCode
(www.codoncode.com) and a consensus sequence was generated in Bioedit 7.0
(www.mbio.ncsu.edu/bioedit/bioedit.html/). Gaps (but not indels) between de novo
contigs were replaced with FvRefv2 sequence forming a chimeric sequence which was
then used for reference guided assembly (RGA) (Shen and Mockler, in prep;
http://rga.cgrb.oregonstate.edu). The resulting ‘chimeric reference sequence’ was
different for each species. The settings used for RGA were ≤2 mismatches per microread,
Q-values ≥20, read depth ≥5, error rate of 0.06 and SNP acceptance requiring ≥70% of
reads to be in agreement.
The online software MAFFT (http://align.bmr.kyushu-u.ac.jp/mafft/software/) was
used to align the assemblies from the RGA output. The MAFFT alignment (134,000 bp)
was manually checked for misalignments and sequencing errors. The aligned sequences
were searched for the presence of primer sequences, and their positions and amplicon
sizes in Fragaria were noted. Nucleotide positions corresponding to primer positions
were eliminated from the alignment because most of the primers used for PCR were not
of Fragaria origin. Indels were scored for future phylogenetic analysis and summary
statistics related to the number of reads in each amplicon were manually calculated.
RESULTS AND DISCUSSION
The 63 primer pairs used in this study were designed to amplify fragments ranging
in size from 871-5317 bp with an average size of 3044 bp. Their genome positions and
sizes were based on the completely sequenced chloroplast genome of Nicotiana tabacum
L. (Shinozaki et al., 1986). Primer sequence positions in FvRefv2 allowed the calculation
of amplicon sizes in Fragaria. Based on these calculations, the Fragaria chloroplast
genome is estimated to be 131,860 bp, which is smaller than that of N. tabacum (155,844
bp). Based on FvRefv2, the mean amplicon size was 3045 bp, median size was 3014 bp,
and the size of amplified fragments ranged from 845-5386 bp. Of the 126 primers used to
317
amplify the Fragaria and Potentilla chloroplast genomes, 20 sequences (forward and/or
reverse) were not found in FvRefv2. The missing primer sequences are attributed to
sequencing failures, the smaller size of the chloroplast genome of Fragaria, and
mutations in primer target regions. Since PCR success was verified by agarose gel
electrophoresis before sequencing, amplified products that were not detected in the
sequenced chloroplasts could represent non-target amplification of nuclear and/or
mitochondrial genome regions.
The median number of reads obtained for each amplicon in the different species
ranged from 0-56 while the average was 16. Amplicons with median numbers ≤5 may
indicate low quantity in the PCR amplicon pool for the species due to underestimation of
PCR product quantity or amplification of non-target DNA. Amplicons with average
median values of <5 reads per base pair corresponded to those flanked by missing primer
sequences. This further indicates non-target amplification of some regions of the
chloroplast genome. Manual improvement of the MAFFT alignment by removing regions
of possible sequencing errors, primer sequences, non-target amplicon sequences, and the
addition of insertions and deletion reduced the length of the aligned sequence from
134,000 to 96,390 bp. Forty-eight indels were noted in the aligned chloroplast sequences.
The final alignment of sequenced genomes was 96,438 bp including indels, and contained
995 variable sites and 319 parsimony-informative sites. The limited variation we
observed in the chloroplast genome of Fragaria represents less than 1% of the MAFFT
alignment, consistent with the low variability noted in previous studies (Harrison et al.,
1997; Potter et al., 2000). However, the few variable sites observed in this study will
prove useful in resolving relationships among Fragaria species as well as in identifying
variable regions in the chloroplast genome that may be used as DNA barcodes.
Summaries of sequencing run output and sequencing analysis results are provided
(Table 2). The genome coverage obtained with RGA (36-82%) was significantly higher
than that obtained with de novo assembly using Velvet Assembler (version 0.4) (7-74%)
(p value=0.00001). RGA aligns microreads to their best match on a reference genome and
thus higher genome coverage was expected. The genome coverage based on RGA results
for three of the species, F. bucharica (69%), F. moschata (46%), and F. nubicola (36%)
was significantly lower than for the remaining species (p value=0.04). The low coverage
observed in these three species could be due to errors during PCR product pooling or in
preparation for sequencing.
CONCLUSIONS
Multiplex sequencing of Fragaria chloroplast genomes resulted in genome
coverages ranging from 73-82% in 16 species, and low genome coverage in F. bucharica,
F. moschata and F. nubicola. To reduce non-specific and null amplifications and errors in
amplicon pooling concentrations, the samples with low coverage will be sequenced
directly using genomic DNA preps as will four additional samples including F. iinumae,
F. pentaphylla, F. chiloensis and F. moupinensis that had genome coverage values ≤76%
from RGA. This may help to complete the dataset enabling more comprehensive
phylogenetic analysis. A preliminary phylogenetic analysis (not shown) supported the
three diploid clades of Rousseau-Gueutin et al. (2009) with bootstrap values ranging from
98-100%. The majority of the remaining nodes were supported by 100% bootstrap values,
which are higher than those in previous Fragaria phylogenetic studies using molecular
markers and nuclear and chloroplast genome sequences (Harrison et al., 1997; Potter et
al., 2000; Rousseau-Gueutin et al., 2009). The results of this study will be useful in
identifying chloroplast genome regions of high sequence variation for testing
evolutionary relationships in future studies such as DNA barcoding.
ACKNOWLEDGMENTS
The authors thank Theodore Bunch, Michael Dossett, April Nyberg and Sarah
Sundholm for invaluable technical support. We also thank Matthew Parks for assistance
with advice in setting up experiments, Dr. Brian Knaus for laboratory and data analysis
318
assistance and, Mark Dasenko and Chris Sullivan (OSU, Center for Genome Research
and Biocomputing) for sequencing and data management support.
Literature Cited
Bringhurst, R.S. 1990. Cytogenetics and evolution in American Fragaria. HortScience
25:879-881.
Bringhurst, R.S. and Voth, V. 1984. Breeding octoploid strawberries. Iowa State Journal
of Research 58:371-381.
Chung, S.M., Gordon, V.S. and Staub, J.E. 2007. Sequencing cucumber (Cucumis sativus
L.) chloroplast genomes identifies differences between chilling-tolerant and ­
susceptible cucumber lines. Genome 50:215-225.
Cronn, R., Liston, A., Parks, M., Gernandt, D.S., Shen, R. and Mockler, T. 2008.
Multiplex sequencing of plant chloroplast genomes using Solexa sequencing-by­
synthesis technology. Nucleic Acids Res. 36:122.
Davis, T.M. and DiMeglio, L.M. 2004. Identification of putative diploid genome donors
to the octoploid cultivated strawberry, Fragaria ×ananassa. Plant and Animal
Genome XII. San Diego, CA, January 10-14. (poster #603).
Harrison, R.E., Luby, J.J. and Furnier, G.R. 1997. Chloroplast DNA restriction fragment
variation among strawberry (Fragaria spp.) taxa. J. Amer. Soc. Hort. Sci. 122:63-68.
Hummer, K. and Bassil, N. 2008. Unexpected polyploidy in wild Asian strawberries
HortScience 43:1187.
Hummer, K., Nathewet, P. and Yanagi, T. 2009. Decaploidy in Fragaria iturupensis
Staudt (Rosaceae). Amer. J. Bot. 96:713-716.
Lawrence, F.J., Galleta, G.J. and Scott, D.H. 1990. Strawberry breeding work of the
United States Department of Agriculture. HortScience 25:895-896.
Nishikawa, T., Salomon, B., Komatsuda, T. and von Bothmer, R. 2002. Molecular
phylogeny of the genus Hordeum using three chloroplast DNA sequences. Genome.
45:1157-1166.
Njuguna, W. and Bassil, N. 2008. A microsatellite fingerprinting set for strawberry,
Fragaria L. HortScience 43: 915.
Potter, D., Eriksson, T., Evans, R.C., Oh, S., Smedmark, J.E.E., Morgan, D.R., Kerr, M.,
Robertson, K.R., Arsenault, M., Dickinson, T.A. and Campbell, C.S. 2007. Phylogeny
and classification of Rosaceae. Plant Syst. Evol. 266:5-43.
Potter, D., Luby, J.J. and Harrison, R.E. 2000. Phylogenetic relationships among species
of Fragaria (Rosaceae) inferred from non-coding nuclear and chloroplast DNA
sequences. Syst. Bot. 25:337-348.
Ravi, V., Khurana, J., Tyagi, A. and Khurana, P. 2006. The chloroplast genome of
mulberry: complete nucleotide sequence, gene organization and comparative analysis.
Tree Gen. Genom. 3:49-59.
Rousseau-Gueutin, M., Gaston, A., Aïnouche, A., Aïnouche, M.L., Olbricht, K., Staudt,
G., Richard, L. and Denoyes-Rothan, B. 2009. Tracking the evolutionary history of
polyploidy in Fragaria L. (strawberry): New insights from phylogenetic analyses of
low-copy nuclear genes. Mol. Phylogen. Evol. 51:515-530.
Senanayake, Y.D.A. and Bringhurst, R.S. 1967. Origin of Fragaria polyploids. I.
Cytological analysis. Amer. J. Bot. 51:221-228.
Staudt, G. 1989. The species of Fragaria, their taxonomy and geographical distribution.
Acta Hort. 265:24-31.
Staudt, G. 2008. Strawberry biogeography, genetics and systematics. In: VI International
symposium, 3-7 March 2008, Huelva.
Staudt, G. and Olbricht, K. 2008. Notes on Asiatic Fragaria species V: F. nipponica and
F. iturupensis. Botanische Jahrbücher für Systematik 127:317-341.
Zerbino, D. and Birney, E. 2008. Velvet: algorithms for de novo short read assembly
using De Bruijn graphs. Genome Res. 18:821-829.
319
Tables
Table 1. Taxa, PI numbers, ploidy and origin of Fragaria accessions used for chloroplast
sequencing.
Taxon
F. orientalis
F. iinumae
F. nipponica
F. virginiana
F. iturupensis
F. viridis
F. ×bifera
F. nilgerrensis
F. bucharica
F. vesca
F. moschata
F. daltoniana
F. chiloensis
F. tibetica
F. pentaphylla
F. gracilis
F. corymbosa
F. ×ananassa ssp. cuneifolia
F. chinensis
F. mandschurica
F. nubicola
320
PI number
PI 551864
PI 637963
PI 637975
PI 612492
PI 641091
PI 616857
PI 616613
PI 616672
PI 551851
PI 551507
PI 551528
PI 641195
PI 612318
PI 651567
PI 651568
CFRA 1908
CFRA 1911
PI 551805
PI 616583
CFRA 1947
PI551853
Ploidy
4x
2x
2x
8x
10x
2x
2x
2x
2x
2x
6x
2x
8x
4x
2x
4x
4x
8x
2x
2x
2x
Origin
Russian Federation
Japan
Japan
Canada
Russian Federation
Sweden
France
China
Pakistan
Germany
France
China
Ecuador
China
China
China
China
United States
China
Mongolia
Pakistan
Table 2. Illumina sequencing output and data summary, including the tag (3 bp barcode used for each sample), the multiplex pool for each
sample, the sum of microreads, average and median number of reads for each base pair position, de novo and reference guided
assembly (RGA) coverage.
Taxon
Tag
Lane/pool
Microreads
F. iinumae
F. corymbosa
F. tibetica
F. viridis
F. daltoniana
P. villosa
F. nipponica
F. bucharica
F. moschata
F. iturupensis
F. nubicola
F.×ananassa ssp. cuneifolia
F. gracilis
F. virginiana
F. pentaphylla
F. chiloensis
F.×bifera
F. mandschurica
F. chinensis
F. vesca
F. orientalis
F. nilgerrensis
acg
agc
ccc
ctg
gta
tac
gat
tca
ggg
tgc
tgc
aac
atg
cac
gct
ttg
acg
agc
ccc
ctg
gta
tac
FragA
607,838
1,600,377
1,193,513
962,316
621,757
1,133,877
579,426
295,691
871,489
1,921,668
1,209,936
864,943
305,281
727,536
757,266
348,758
470,616
682,461
820,618
647,651
866,894
1,042,414
FragB
FragC
FragD
Average
reads/bp
51.62
205.47
161.60
113.77
80.72
112.33
47.95
7.37
72.70
121.96
92.35
108.11
28.00
79.40
21.93
19.45
43.80
72.96
107.22
87.51
97.79
137.55
Median
reads/bp
6
50
14
19
15
4
10
2
0
13
0
12
8
14
4
3
7
10
17
8
14
13
De novo coverage
(%)
43
74
65
68
64
54
58
10
10
61
7
64
51
63
27
22
48
55
69
51
63
62
RGA coverage
(%)
76
82
80
80
81
73
80
69
46
78
36
82
79
79
75
73
79
81
81
79
79
80
321
321
322 
Download